Accudynetest logo

Products available online direct from the manufacturer

ACCU DYNE TEST ™ Bibliography

Provided as an information service by Diversified Enterprises.

3040 results returned
showing result page 36 of 76, ordered by
 

432. Cai, G., M.H. Litt, and I.M. Krieger, “Surface properties and abhesion of undecyl oxazoline block and homopolymers,” J. Polymer Science Part B: Polymer Physics, 29, 773-784, (1991).

The surface properties of three undecyl oxazoline homopolymers and two phenyl/undecyl oxazoline block copolymers (as comparison) were studied. After coating on glass slides and annealing, all films had a low critical surface energy of 21 dynes/cm. Water contact angles were higher than 107° for the most hydrophobic films. The deduction that the polymer surfaces contained close-packed methyl groups was further confirmed by electron spectroscopy chemical analysis (ESCA) angle profiling on an annealed undecyl oxazoline homopolymer film. A model was developed for the variation of elemental ratios as a function of photoelectron take-off angle. This verified that the polymer films had the polymer backbones parallel to the surface with the undecyl tails oriented toward the surface. When these block and homopolymers were coated on copy paper and glass slides, the peel strengths of pressure-sensitive adhesives with these surfaces were very low for short dwell times at room temperature. At long dwell times or at elevated temperatures, the peel strengths remained low for the homopolymers but increased greatly for the block copolymers to values higher than those in the tape on glass. After 24 h at 70°C, ESCA analysis showed that the adhesive diffused into the phenyl block domains of the diblock copolymer, generating high peel strength and cohesive failure. However, under the same annealing conditions, the triblock copolymer showed adhesive failure while peel strength increased. ESCA analysis showed very litle diffusion of the adhesive into the triblock copolymer. The homopolymers were stable toward vinyl acetate type adhesives even at elevated temperature; they were abhesive up to 100°C with no interdiffusion.

604. Yasuda, T., K. Yoshida, T. Okuno, and H.K. Yasuda, “A study of surface dynamics of polymers, III. Surface dynamic stabilization by plasma polymerization,” J. Polymer Science Part B: Polymer Physics, 26, 2061-2074, (1988).

As demonstrated in Part II of this series of studies, the hydrophobic character of CF4 plasma-treated Nylon 6 and poly(ethylene terephthalate) (PET) decay with time of water immersion, and the rate of decay can be used as a measure for the surface mobility of (substrate) polymers. The same method of using fluorine-containing moieties introduced by CF4 plasma treatment as surface labeling is applied to investigate the influence of a thin layer of plasma polymer of methane applied onto the surface of those polymers. An ultrathin layer of plasma polymer provides a barrier to the rotational and diffusional migration of the introduced chemical moieties from the surface into the bulk of the film. The influence of operational parameters of plasma polymerization on the surface dynamic stability are examined by measuring the decay rate constants for (subsequently) CF4 plasma-treated samples. The rate constant was found to decrease sharply with increasing value of plasma energy input manifested by J/kg monomer, and no decay was observed as the energy input reached a threshold value (about 6.5 GJ/kg for PET, about 7.0 GJ/kg for Nylon 6), indicating that unperturbable surfaces can be created by means of plasma polymerization.

816. Sanchis, R.M., O. Calvo, L. Sanchez, D. Garcia, and R. Balart, “Enhancement of wettability in low density polyethylene films using low pressure glow discharge N2 plasma,” J. Polymer Science Part B: Polymer Physics, 45, 2390-2399, (2007).

Low pressure glow discharge nitrogen plasma has been used to improve wettability in a low density polyethylene (LDPE) film for technical applications. The plasma treatment was carried out at a power of 300 W for different exposure times in the 1–20 min range. Wettability changes were analyzed using contact angle measurements. In addition to this, plasma-treated samples were subjected to an aging process to determine the durability of the plasma treatment. X-ray photoelectron spectroscopy, atomic force microscopy, and scanning electron microscopy were used for surface characterization. The nitrogen plasma treatment considerably reduced contact angle values thus indicating an increase in surface wettability. The spectroscopic study showed presence of oxygen-based species on the plasma-treated samples, which are mainly generated after the plasma treatment as a consequence of air exposure. These polar species contribute to improve surface functionalization, but this is almost lost during aging due to the hydrophobic recovery process. Microscopic studies revealed that also small changes in surface roughness occurred during the plasma treatment but these are very low compared to surface activation. The results confirmed that low pressure nitrogen can be considered as an environmentally efficient process to improve wettability in low density polyethylene films. © 2007 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys 45: 2390–2399, 2007

1283. Zheng, Z., X. Wang, M. Shi, and G. Zhou, “Surface modification of ultrahigh-molecular-weight polyethylene fibers,” J. Polymer Science Part B: Polymer Physics, 42, 463-472, (Feb 2004).

To prevent the loss of fiber strength, ultrahigh-molecular-weight polyethylene (UHMWPE) fibers were treated with an ultraviolet radiation technique combined with a corona-discharge treatment. The physical and chemical changes in the fiber surface were examined with scanning electron microscopy and Fourier transform infrared/attenuated total reflectance. The gel contents of the fibers were measured by a standard device. The mechanical properties of the treated fibers and the interfacial adhesion properties of UHMWPE-fiber-reinforced vinyl ester resin composites were investigated with tensile testing. After 20 min or so of ultraviolet radiation based on 6-kW corona treatment, the T-peel strength of the treated UHMWPE-fiber composite was one to two times greater than that of the as-received UHMWPE-fiber composite, whereas the tensile strength of the treated UHMWPE fibers was still up to 3.5 GPa. The integrated mechanical properties of the treated UHMWPE fibers were also optimum. © 2003 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys 42: 463–472, 2004
https://onlinelibrary.wiley.com/doi/10.1002/polb.10727

1320. Kwok, D.Y., A. Li, and A.W. Neumann, “Low-rate dynamic contact angles on poly(methyl methacrylate/ethyl methacrylate, 30/70) and the determination of solid surface tensions,” J. Polymer Science Part B: Polymer Physics, 37, 2039-2051, (1999).

Low-rate dynamic contact angles of 12 liquids on a poly(methyl methacrylate/ethyl methacrylate, 30/70) P(MMA/EMA, 30/70) copolymer were measured by an automated axisymmetric drop shape analysis-profile (ADSA-P). It was found that five liquids yield nonconstant contact angles, and/or dissolve the polymer on contact. From the experimental contact angles of the remaining seven liquids, it is found that the liquid–vapor surface tension times cosine of the contact angle changes smoothly with the liquid–vapor surface tension (i.e., γl|Kv cos θ depends only on γl|Kv for a given solid surface or solid surface tension). This contact angle pattern is in harmony with those from other methacrylate polymer surfaces previously studied.45,50 The solid–vapor surface tension calculated from the equation-of-state approach for solid–liquid interfacial tensions14 is found to be 35.1 mJ/m2, with a 95% confidence limit of ± 0.3 mJ/m2, from the experimental contact angles of the seven liquids. © 1999 John Wiley & Sons, Inc. J Polym Sci B: Polym Phys 37: 2039–2051, 1999
https://onlinelibrary.wiley.com/doi/abs/10.1002/(SICI)1099-0488(19990815)37:16%3C2039::AID-POLB8%3E3.0.CO;2-O

1840. Schonhorn, H., and L.H. Sharpe, “Surface tension of molten polypropylene,” J. Polymer Science Part B: Polymer Physics, 3, 235-237, (1965).

2498. Aouinti, M., A. Gibaud, D. Chateigner, and F. Poncin-Epaillard, “Morphology of polypropylene films treated in CO2 plasma,” J. Polymer Science Part B: Polymer Physics, 42, 2007-2013, (May 2004).

One of the most important claims for the plasma technique as a surface treatment is that it modifies only a few atomic layers of materials. However, with polymers, this assumption must be carefully verified to keep the bulk mechanical properties constant. Besides the oxidation of the film, with specific plasma conditions such as high power and duration, the polypropylene film structure is also modified in the bulk through vacuum ultraviolet absorption and thermal relaxation. This change is associated with smectic- and amorphous-phase transformation into an α-monoclinic phase, with a rapid rate for the smectic transformation and a slower rate for the amorphous transformation. At the same time, the crystallite size increases, and the polypropylene film texture is planar and moderated (1.7 mrd at the maximum of the distribution, with a discharge power of 100 W and a treatment duration of 10 min). © 2004 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys 42: 2007–2013, 2004
https://onlinelibrary.wiley.com/doi/abs/10.1002/polb.20071

2516. Inagaki, N., K. Narushima, N. Tuchida, and K. Miyazaki, “Surface characterization of plasma-modified poly(ethylene terephthalate) film surfaces,” J. Polymer Science Part B: Polymer Physics, 42, 3727-3740, (Oct 2004).

Poly(ethylene terephthalate) (PET) film surfaces were modified by argon (Ar), oxygen (O2), hydrogen (H2), nitrogen (N2), and ammonia (NH3) plasmas, and the plasma-modified PET surfaces were investigated with scanning probe microscopy, contact-angle measurements, and X-ray photoelectron spectroscopy to characterize the surfaces. The exposure of the PET film surfaces to the plasmas led to the etching process on the surfaces and to changes in the topography of the surfaces. The etching rate and surface roughness were closely related to what kind of plasma was used and how high the radio frequency (RF) power was that was input into the plasmas. The etching rate was in the order of O2 plasma > H2 plasma > N2 plasma > Ar plasma > NH3 plasma, and the surface roughness was in the order of NH3 plasma > N2 plasma > H2 plasma > Ar plasma > O2 plasma. Heavy etching reactions did not always lead to large increases in the surface roughness. The plasmas also led to changes in the surface properties of the PET surfaces from hydrophobic to hydrophilic; and the contact angle of water on the surfaces decreased. Modification reactions occurring on the PET surfaces depended on what plasma had been used for the modification. The O2, Ar, H2, and N2 plasmas modified mainly CH2 or phenyl rings rather than ester groups in the PET polymer chains to form CSingle BondO groups. On the other hand, the NH3 plasma modified ester groups to form CSingle BondO groups. Aging effects of the plasma-modified PET film surfaces continued as long as 15 days after the modification was finished. The aging effects were related to the movement of CDouble BondO groups in ester residues toward the topmost layer and to the movement of CSingle BondO groups away from the topmost layer. Such movement of the CDouble BondO groups could occur within at least 3 nm from the surface. © 2004 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys 42: 3727–3740, 2004
https://onlinelibrary.wiley.com/doi/abs/10.1002/polb.20234

2427. Kim, C.Y., G. Suranyi, and D.A.I. Goring, “Corona induced bonding of synthetic polymers to cellulose,” J. Polymer Science Part C: Polymer Symposia, 30, 533-542, (1970).

Corona treatment improved bonding between sheets of cellulose and synthetic polymers. The bond strength increased at higher temperatures of pressing. Physical changes in the surface were detected microscopically after corona treatment in air. Sheets treated in pure nitrogen made strong bonds although the surface treated in nitrogen was indistinguishable from the untreated surface.

1818. Pittman, A.G., D.L. Sharp, and B.A. Ludwig, “Polymers derived from fluoroketones II: Wetting properties of fluoroalkyl acrylates and methacrylates,” J. Polymer Science, Part A-1: Polymer Chemistry, 6, 1729-1740, (1968).

The critical surface tension of wetting (γc) for certain branched-chain polymeric fluoroalkyl acrylates and methacrylates was obtained. Polymeric materials utilized in this study can be represented by the repeating units

mathmatical formual
, where R is H or CH3, R′ is H or F, and X is F or Cl, by mathmatical formual, where n is 2, 5, or 11, and by mathmatical formual, where R is H or CH3 and n′ is 2 or 6. Monomer synthesis involved either the direct acylation of a fluoroketone–metal fluoride adduct or a fluoroalcohol with acryloyl or methacryloyl chloride or a displacement reaction between a fluoroketone–metal fluoride adduct and an ω-bromoester. In general, modifications in the pendent fluoroalkyl group affected γc in a manner predictable from previous work by Zisman et al.; e.g., γc was increased when either H or Cl was substituted for F in the side chain. In polymeric alkyl acrylates containing a heptafluoroisopropyl side chain γc increased as the fluorocarbon group was removed from the proximity of the polymer backbone by intervening methylene groups. A comparison of the wetting properties of polyacrylates containing either a perfluoroisopropyl or n-perfluoropropyl group showed that the polymer containing the isopropyl group had a lower γc.

1849. Clark, D.T., and A. Dilks, “ESCA applied to polymers, XXIII: RF glow discharge modification of polymers in pure oxygen and helium-oxygen mixtures,” J. Polymer Science, Part A: Polymer Chemistry, 17, 957-976, (1979).

The oxidation of polyethylene, polypropylene, and polystyrene by exposure to plasmas excited in pure oxygen and helium–oxygen mixtures at low power levels has been studied. A detailed curve resolution procedure is outlined, and the rate of oxidation is shown to be a strong function of the polymer structure for pure oxygen plasmas, as is the composition of the oxidized layer; this is not the case, however, for oxidation effected by helium–oxygen mixtures. It seems likely, from a consideration of the available data, that the oxidation is confined to the outermost monolayer and is initiated by a crosslinking mechanism that involves oxygen-containing functionalities.

155. Hansen, R.H., and H. Schonhorn, “A new technique for preparing low surface energy polymers for adhesive bonding,” J. Polymer Science, Polymer Letters Edition, 4, 203-209, (1966).

Contact time of activated gas with polymer film of as little as 1 sec. under these relatively mild conditions resulted in greatly improved adhesive joint strength for polyethylene. Longer contact times were required for polymers such as polytetrafluoroethylene. Helium, argon, krypton, neon, and xenon, and even hydrogen and nitrogen were all effective crosslinking agents although the latter also changed wettability of the surface. Adhesive joints were prepared by sandwiching 1041 films of polyethylene (Marlex 5003, Phillips Petroleum Co., Bartlesville, Oklahoma) and polytetrafluoroethylene (G-80, Allied Chemical Co., Morristown, New Jersey), before and after CASING, between epoxy-coated aluminum strips (3). The values obtained for tensile shear strengths of these joints are shown in Figure 1. In these instances, the polyethylene film was treated for 10 sec. and the polytetrafluoroethylene film was treated for 10 min.

2428. Courval, G.J., D.G. Gray, and D.A.I. Goring, “Chemical modification of polyethylene surfaces in a nitrogen corona,” J. Polymer Science: Polymer Letters Edition, 14, 231-235, (Apr 1976).

1692. Strom, G., “The importance of surface energetics and dynamic wetting in offset printing,” J. Pulp and Paper Science, 19, J79, (1993).

The surface energetic properties of different areas of the offset printing plate are the key factors of this printing process, since they control the ink transfer during printing. The importance of these factors is discussed for both waterless offset and conventional offset. The printing process is highly dynamic. New surfaces are created and their lifetimes are short. From recent theories of dynamic wetting, it has been concluded that spontaneous removal of ink films from nonimage areas is a very slow due to the high ink viscosity and the low dynamic contact angle. Thus it is of less importance.

2783. Aspler, J.S., S. Davis, and M.B. Lyne, “The surface chemistry of paper in relation to dynamic wetting and sorption of water and lithographic fountain soutions,” J. Pulp and Paper Science, 13, 355-360, (1987).

1341. Padday, J.F., “Apparatus for measuring the spreading coefficient of a liquid, on a solid surface,” J. Scientific Instrumentation, 36, (1959).

2737. Mukhopadhyay, S., and R. Fangueiro, “Physical modification of natural fibers and thermoplastic films for composites - a review,” J. Thermoplastic Composite Materials, 22, 135-162, (Mar 2009).

The article throws light on the physical methods to modify natural fibers to be used in composites. Physical methods in natural fiber processing are used to separate natural fiber bundles into individual filaments and to modify the surface structure of the fibers so as to improve the use of natural fibers in composites. Steam explosion and thermomechanical processes fall in the first category while plasma, dielectric barrier techniques and corona fall in the second. The physical treatments have also been used to modify the thermoplastic polymeric films like polyethylene and polypropylene in a bid to impart reactivity. Reviewing such developments, the areas for further research are suggested.

31. Bodo, P., and J.-E. Sundgren, “Adhesion of evaporated titanium to polyethylene: effects of ion bombardment pretreatment,” J. Vacuum Science and Technology, A2, 1498-1502, (1984).

Titanium films, 1 μm thick were electron‐beam evaporated onto polyethylene (PE) that had been pretreated in situ by 2 keV Ar+ bombardment. A measure of the film adhesion was obtained by measuring the pull strength required to remove the Ti films. A strong dependence of the adhesion on the ion dose was found. The pull strength had a maximum of approximately 20 MPa after a dose of 6×1014 ions/cm2 but decreased for higher ion doses. Without any ion bombardment prior to deposition, the adhesion was very poor with a pull strength of approximately 2 MPa. XPS analysis was used to examine the effect of the ion bombardment on the chemistry of the PE substrate and the Ti/PE interface. Untreated PE samples were contaminated with surface impurities and probably also with low molecular weight hydrocarbons. As the adhesion is maximized, most of the impurities are removed by the ion bombardment. The strong adhesion is suggested to be due to formation of a carbidelike Ti–C interfacial layer, detected by XPS.

234. Matienzo, L.J., F. Emmi, F.D. Egitto, et al, “Surface composition and distribution of fluorine in plasma-fluorinated polyimide,” J. Vacuum Science and Technology, A6, 950-953, (1988).

Surface composition, fluorine distribution, and morphology were determined for polyimide films modified downstream from microwave plasmas containing CF4/O2. Complementary analytical techniques including x‐ray photoelectron spectroscopy, Rutherford backscattering spectroscopy, and scanning electron microscopy yielded a more complete understanding of polyimide fluorination and subsequent etching of the modified film. Depth of fluorination increased nonlinearly with treatment time for films exposed downstream from a CF4‐rich plasma. Exposure downstream from an O2‐rich plasma resulted in a reduction of thickness in both the fluorinated layer and the unmodified polyimide during etching. Finally, a model for fluorination of polyimide and subsequent removal is proposed.

466. Grant, J.L., D.S. Dunn, and D.J. McClure, “Argon and oxygen sputter etching of polystyrene, polypropylene, and poly(ethylene terephthalate) thin films,” J. Vacuum Science and Technology, A6, 2213-2220, (1988).

Surface chemical modification of polymer thin films induced by sputter etching was studied by x‐ray photoelectron spectroscopy (XPS) and infrared reflection–absorption spectroscopy (IRRAS). The polymers studied were polystyrene, polypropylene, and poly(ethylene terephthalate) (PET). Oxygen and argon sputter etching of these polymers causes surface oxidation and possibly crosslinking; trends in polymer oxidation can be correlated with the etchant gas, etch power, and initial material properties. For polystyrene and polypropylene, the predominant new functionalities formed are CDouble BondO and CSingle BondO groups; the breadth of the infrared absorption bands suggests that many different types of these groups exist. For PET, the predominant damage mechanism is crosslinking, with only a slight degree of oxidation resulting from oxygen sputter etching. This work suggests that the information provided by XPS and IRRAS is highly complimentary and will be useful in future studies of polymer functionalization and derivatization.

567. Sengupta, K.S., and H.K. Birnbaum, “Structural and chemical effects of low-energy ion bombardment of PMMA-ODA surfaces,” J. Vacuum Science and Technology, A9, 2928-2935, (1991).

The effects of ion irradiation on polyimide surfaces have been studied using x‐ray photoemission techniques. Ion bombardment with energies in the range 0.5–2.0 keV and doses between 8×1013 and 1×1015 ions/cm2 were carried out in situ in the x‐ray photoelectron spectrometer and the chemistry of the modified surface was monitored using core level spectral changes. At low doses and energies, carbonyl groups were preferentially sputtered keeping the rest of the monomer intact. Loss of nitrogen was insignificant compared to losses of carbon and oxygen. At higher energies and doses, the polymer undergoes extensive bond scission, restructuring of various functional groups and species, together with radical and anion formation. High resolution spectra indicated a binding energy scale shift to a lower value, which increased with ion energy and dose, and which was related to the creation of a surface negative charge. The effects of exposure to moisture in the ambient on the surface charge, on the surface structure, and on the surface chemistry was studied.

1309. Callen, B.W., M.L. Ridge, S. Lahooti, A.W. Neumann, and R.N.S. Sodhi, “Remote plasma and UV-ozone modification of polystyrene,” J. Vacuum Science and Technology, A13, 2023-2029, (1995).

Remote argon plasma (RP) and ozone in the presence of ultraviolet light (UV–O3) were used to render polystyrene (PS) surfaces hydrophilic in a controlled manner for eventual application in cell‐surface interaction studies. X‐ray photoelectron spectroscopy (XPS) was used to characterize both methods of modification. The degree of modification on PS was measured by an increase in surface oxygen and concomitant change in C 1s binding energies as a function of time. Both remote plasma and UV–O3 are shown to be partially surface destructive, producing polymer fragments which are easily washed away to leave stable modified surfaces of oxidized polymer comprising of distributions of C–O, C=O and O—C=O type groups. Of the two methods, UV–O3 is shown to be more versatile and conducive to preparing PS surfaces with controllably varying degrees of modification. UV–O3 modified polystyrene is shown to be stable in air for at least eight months. Contact angle methods were used in correlation with XPS in characterizing UV–O3 modified surfaces. It is shown that changes in surface tension and total surface oxygen content were related, however, not directly connected.

1736. Kumagai, H., H. Denbo, N. Fujii, and T. Kobayashi, “Poly(ethylene terephthalate) decomposition process in oxygen plasma: Emission spectroscopic and surface analysis for oxygen-plasma reaction,” J. Vacuum Science and Technology, A22, 1-7, (2004).

Emission spectroscopy was applied to observe the reaction process of poly (ethylene terephthalate) (PET) in an oxygen (O2) plasma generated by a microwave discharge. As the PET was exposed in the O2 plasma flow, light emitted from the PET surface was monitored. In the diagnosis measurement, several emission peaks assigned to the Hα atomic line at 652 nm, Hβ at 486 nm, OH (2Σ→2Π) transition near 244–343 nm and CO (b3 Σ→a3 Σ) near 283–370 nm were observed and measured at various discharge times. These results indicated that after the plasma etching, the PET sample was decomposed by the oxygen plasma reaction, and then, hydrogen abstraction and carbon oxidation processes. We also observed the time profile of oxygen atom, as the atom-emission intensity at 777 nm was monitored. As Hβ atomic and OH molecule lines appeared in the presence of PET, the O atom intensity was significantly reduced. In the surface analysis on Fourier transform infrared and x-ray photoelectron spectroscopy measurements, it was found that for the PET surface treated by O2 plasma containing excited atomic oxygen species, ester bands were broken and carbonization formed on the PET surface.

2070. Hozumi, A., N. Shirahata, Y. Nakanishi, S. Asakura, and A. Fuwa, “Wettability control of a polymer surface through 126 nm vacuum ultraviolet light irradiation,” J. Vacuum Science and Technology, A22, 1309-1314, (Jul 2004).

The control of the surface wettability of poly (methyl methacrylate) (PMMA) substrates has been successfully demonstrated using an Ar2* excimer lamp radiating 126 nm vacuum ultraviolet (VUV) light. Each of the samples was exposed to 126 nm VUV light in air over the pressure range of 2×10−4-105 Pa. Although at the process pressures of 10, 103, and 105 Pa, the PMMA surfaces became relatively hydrophilic, the degree of hydrophilicity depended markedly on the pressure. The minimum water contact angles of the samples treated at 10, 103, and 105 Pa were about 50°, 33°, and 64°, respectively. These values were larger than those of PMMA substrates hydrophilized through 172 nm VUV irradiation conducted under the same conditions. On the other hand, after 126 nm VUV irradiation conducted under the high vacuum condition of 2×10−4 Pa, the PMMA substrate surface became carbon-rich, probably due to preferential cross-linking reactions, as evidenced by x-ray photoelectron spectroscopy. This surface was hydrophobic, showing a water contact angle of about 101°. Although the 126 nm VUV-irradiated surfaces appeared relatively smooth when observed by atomic force microscope, very small particles with diameters of 30-60 nm, which probably originated from the readhesion of photodecomposed products, existed on all of the sample surfaces.

2084. Lee, Y., S. Han, J.-H. Lee, J.-H. Yoon, H.E. Lim, and K.-J. Kim, “Surface studies of plasma source ion implantation treated polystyrene,” J. Vacuum Science and Technology, A16, 1710-1715, (May 1998).

The plasma source ion implantation (PSII) was utilized to improve the wettability and the stability of surface layer formed in the modification of polymeric materials. Polystyrene was treated with different kinds of plasma ions to render the surface more hydrophilic or hydrophobic. Hydrophobic recovery of PSII-treated polystyrene was also observed as a function of aging time, aging temperature, and treatment parameters. Treatment parameters involve kinds of gases, pressure, plasma power, pulse frequency, pulse voltage, etc. To study the effect of inert gas on hydrophobic recovery, polystyrene samples were prepared by helium, argon, or gas-mixture treatment. Time-of-flight secondary ion mass spectrometry (TOF-SIMS) has been used to interpret the PSII-treated polystyrene surface and its hydrophobic recovery, with the assistance of x-ray photoelectron spectroscopy and water contact angle measurements. TOF-SIMS spectra of O218 PSII-treated samples showed the presence of O18-containing peaks from the modified surfaces. PSII modifications provide more stable surfaces of polystyrene as a function of aging time than plasma treatments. The comparison of aging behavior data allowed for examination of the differences in the stability of the functionality introduced by the two different treatment techniques.

3029. O'Kane, D.F., and K.L. Mittal, “Plasma cleaning of metal surfaces,” J. Vacuum Science and Technology, 11, 567-569, (1974).

Mild plasma cleaning of metal surfaces was shown to be effective in removing organic contaminants. Auger electron spectroscopy and surface wettability measurements were used to evaluate the plasma cleaning procedure and to provide a comparison with conventional solvent cleaning methods.

2022. Matienzo, L.J., J.A. Zimmerman, and F.D. Egitto, “Surface modification of fluoropolymers with vacuum ultraviolet irradiation,” J. Vacuum Science and Technology A, 12, 2662-2671, (Sep 1994).

2031. Lim, H., Y. Lee, S. Han, and J. Cho, “Surface treatment and characterization of PMMA, PHEMA, and PHPMA,” J. Vacuum Science and Technology A, 19, 1490-1496, (Jul 2001).

Poly(methylmethacrylate) (PMMA), poly(2-hydroxyethyl methacrylate (PHEMA), and poly(2-hydroxypropyl methacrylate) (PHPMA) were modified to improve the wettability by two techniques: plasma and plasma source ion implantation. The modified surfaces were characterized to investigate the dependence of the modification and hydrophobic recovery on the polymer structure. The differences obtained under optimal experiment conditions among the polymers were interpreted in terms of their polymer structures including the glass transition temperature. The surface free energy, calculated from the contact angle measurements, revealed that its polar component was a dominant factor in improving the wettability. The PSII treatment created more functional groups on the surface and extensively modified the polymer layer than the plasma treatment.

1085. Cheng, F., S.G. Hong, and C.A. Ho, “The adhesion properties of an ozone modified thermoplastic olefin elastomer,” J.Adhesion, 123-138, (1998) (also in Fundamentals of Adhesion and Interfaces, L.P. DeMejo, D.S. Rimai, and L.H. Sharpe, eds., Jan 2000, Gordon and Breach Science Publ., p.123-138).

812. Murahara, M., and K. Toyoda, “Excimer laser-induced photochemical modification and adhesion improvement of a fluororesin surface,” J.Adhesion Science and Technology, 9, 1601-1609, (1995) (also in Polymer Surface Modification: Relevance to Adhesion, K.L. Mittal, ed., p. 213-222, VSP, May 1996).

1116. Schoff, C.K., “Coatings clinic: Wetting and wettability,” JCT CoatingsTech, 1, 108, (Oct 2004).

1167. Schoff, C.K., “Coatings clinic: Surface tension and surface energy,” JCT CoatingsTech, 3, 72, (Feb 2006).

1566. Williams, K., and B. Bauman, “New technology for enhancing wood-plastic composites,” JCT CoatingsTech, 4, 52-57, (Aug 2007).

2154. Schoff, C.K., “Coatings clinic: Interfaces and migration,” JCT CoatingsTech, 6, 48, (May 2009).

2157. no author cited, “Two new coatings-related standards released by ASTM International,” JCT CoatingsTech, 6, 19, (Jun 2009).

1006. Kusano, Y., T. Noguchi, M. Yoshikawa, N. Kato, and K. Naito, “Effect of discharge treatment on vulcanised rubber surfaces,” in IRC '95 Kobe International Rubber Conference Proceedings, 432-435, Japan Society of Rubber Industry, 1995.

9. Andrews, E.H., and N.E. King, “Surface energetics and adhesion,” in Polymer Surfaces, 47-63, John Wiley & Sons, 1978.

43. Briggs, D., “Analysis and chemical imaging of polymer surfaces by SIMS,” in Polymer Surfaces and Interfaces, Feast, W.J., and H.S. Munro, eds., 33-53, John Wiley & Sons, 1987.

62. Clark, D.T., and W.J. Feast, eds., Polymer Surfaces, John Wiley & Sons, 1978.

71. Davies, M.C., “SSIMS - an emerging technique for the surface chemical analysis of polymeric biomaterials,” in Polymer Surfaces and Interfaces II, Feast, W.J., H.S. Munro, and R.W. Richards, eds., 203-226, John Wiley & Sons, Apr 1993.

 

<-- Previous | 1 | 2 | 3 | 4 | 5 | 6 | 7 | 8 | 9 | 10 | 11 | 12 | 13 | 14 | 15 | 16 | 17 | 18 | 19 | 20 | 21 | 22 | 23 | 24 | 25 | 26 | 27 | 28 | 29 | 30 | 31 | 32 | 33 | 34 | 35 | 36 | 37 | 38 | 39 | 40 | 41 | 42 | 43 | 44 | 45 | 46 | 47 | 48 | 49 | 50 | 51 | 52 | 53 | 54 | 55 | 56 | 57 | 58 | 59 | 60 | 61 | 62 | 63 | 64 | 65 | 66 | 67 | 68 | 69 | 70 | 71 | 72 | 73 | 74 | 75 | 76 | Next-->